Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

minerals

Article
Proposed Methodology to Evaluate CO2 Capture
Using Construction and Demolition Waste
Domingo Martín 1 , Vicente Flores-Alés 2 and Patricia Aparicio 1, *
1 Departamento de Cristalografía, Mineralogía y Q. Agrícola, Facultad de Química, Universidad de Sevilla,
41012 Sevilla, Spain; dmartin5@us.es
2 Departamento de Construcciones Arquitectónicas II, Escuela Técnica Superior de Ingeniería de Edificación,
Universidad de Sevilla, 41012 Sevilla, Spain; vflores@us.es
* Correspondence: paparicio@us.es

Received: 8 August 2019; Accepted: 1 October 2019; Published: 5 October 2019 

Abstract: Since the Industrial Revolution, levels of CO2 in the atmosphere have been constantly
growing, producing an increase in the average global temperature. One of the options for Carbon
Capture and Storage is mineral carbonation. The results of this process of fixing are the safest in
the long term, but the main obstacle for mineral carbonation is the ability to do it economically in
terms of both money and energy cost. The present study outlines a methodological sequence to
evaluate the possibility for the carbonation of ceramic construction waste (brick, concrete, tiles) under
surface conditions for a short period of time. The proposed methodology includes a pre-selection of
samples using the characterization of chemical and mineralogical conditions and in situ carbonation.
The second part of the methodology is the carbonation tests in samples selected at 10 and 1 bar
of pressure. The relative humidity during the reaction was 20 wt %, and the reaction time ranged
from 24 h to 30 days. To show the effectiveness of the proposed methodology, Ca-rich bricks were
used, which are rich in silicates of calcium or magnesium. The results of this study showed that
calcite formation is associated with the partial destruction of Ca silicates, and that carbonation was
proportional to reaction time. The calculated capture efficiency was proportional to the reaction time,
whereas carbonation did not seem to significantly depend on particle size in the studied conditions.
The studies obtained at a low pressure for the total sample were very similar to those obtained for
finer fractions at 10 bars. Presented results highlight the utility of the proposed methodology.

Keywords: suitable methodology for mineral carbonation; construction and demolition waste

1. Introduction
CO2 emissions into the atmosphere are a growing environmental problem in different industrial
sectors. The construction sector is no stranger to this problem considering gaseous emissions are
derived from manufacturing processes of materials used mostly in construction, such as ceramic
materials or cement [1]. Independent of the existing controversy regarding the uncertainty of the
sensitivity of the climate in the international scientific community [2], the most negative forecasts
consider some of the effects predicted by the uncontrolled emission of greenhouse gases as a possible
increase in Earth’s temperature, climatic alterations that will accelerate desertification, and a possible
loss of part of the coastline due to the rise in sea levels.
In the current scenario, the challenges of reducing emissions cannot solely be met with greater
energy efficiency and renewable energy resources in the generation phase. It is absolutely necessary to
act on the management and treatment of emissions. For this reason, research aimed at capturing CO2
is of vital importance in order to achieve the standards set as an objective.

Minerals 2019, 9, 612; doi:10.3390/min9100612 www.mdpi.com/journal/minerals


Minerals 2019, 9, 612 2 of 14

Mineral-carbonation systems for CO2 fixation are another option for Carbon Capture and Storage
(CCS). Although less efficient than geological storage, they are much simpler, cheaper, and have fewer
requirements, which responds to the requirement of ecological rationality raised above. The present
work deals with the possibilities of using construction and demolition waste as CO2 reservoirs.
Preliminary research has shown that construction materials containing calcium and/or magnesium in
their composition in the form of silicates, oxides, and hydroxides, which can react with CO2 to give
rise to carbonates, thus constituting a possible alternative for mineral carbonation from ceramic waste
within options for CO2 capture/storage [3–6]. This possibility is based on the capacity of construction
and demolition waste with a high content of calcium and/or magnesium in the form of silicates, oxides,
and hydroxides to fix CO2 under optimal conditions, using a chemical reaction whose product is the
formation of carbonates and silica as stable by-products. This process of carbonation occurs naturally
with very slow kinetics; it is of interest to design a system and methodology to accelerate this process
to make the capture of CO2 from anthropogenic activity profitable, achieving maximal industrial and
energy efficiency. There are several patents and studies that use residues with a high calcium content
from different types of industrial waste for their carbonation [7–22].
The present work presents a methodological sequence for the control and validation of a viable
alternative of CCS using mineral carbonation of construction and demolition waste (bricks, concrete,
tiles) with a high content of silicates rich in calcium and magnesium. In this case, those with a high
content of ceramic materials and cementitious conglomerate (mortar and concrete) can be considered
optimal landfills for their transformation into CO2 sinks [23,24].
A laboratory model was tested that indicated the efficiency of the system, ensuring extrapolated
conditions at a landfill scale were in accordance with the difficulty involved in the reproduction
of parameters that can be controlled at a laboratory scale. In this sense, it is necessary to establish
a system that approximately reproduces conditions of isolation, humidity, dimensions, and others
that, with generic character, can be reached in a controlled residue deposit. The physical–chemical
and mineralogical mechanisms, the external conditions that have a decisive influence on the process,
and the kinetics of capture and carbonation reactions that favor gas fixation and stabilization were
analyzed. The control and verification procedures of the process were also analyzed, which allowed
the adequate monitoring and optimization of the process.

2. Materials and Methods

2.1. Materials and Carbonation Test


The material used for the development of this methodology proposal was a brick type widely
used in construction, which was specifically a clinker brick (MPC2) from Malpesa S.L. (Bailén, South of
Spain) fired at 1050 ◦ C.
The samples were subjected to a crushing process for granulometric conditioning. Fractions of
less than 4 mm were selected to obtain size ratios consistent with the diameter of the reactor, obtaining
homogeneous distribution so that there was a predominance of coarser fractions corresponding to
conventional waste shredding.
Carbonation tests were carried out in a 0.3 L volume hermetic reactor (Parr Instruments Co.,
Moline, IL, USA). The fixed conditions were 10 bars of CO2 , 4:1 solid–water ratio, and room temperature.
The variable conditions were reaction time (between 24 and 720 h) and particle size (<4, 2–4, and 1–2 mm).
These particle-size fractions were selected because they were the most representative results of the
crushing treatment. Additionally, a test was carried at low pressure (1 bar), room temperature, and a
4:1 solid–water ratio in a 5 L volume hermetic reactor of continuous flow to maintain pressure at 1 bar
during the 720 h of reaction time.
Post-treatment, the samples were dried at 100 ◦ C for 24 h, powdered, and sieved at 50 µm for
subsequent analysis.
Minerals 2019, 9, 612 3 of 14

2.2. Instrumental Techniques: Methodology


Major multielemental chemical composition (in oxides) was performed with an automated
Panalytical Axios model wavelength-dispersive X-ray fluorescence spectrometer (WDXRF). The samples
were prepared for analysis as glass discs to reduce the “matrix effect.”
The mineralogical composition of the untreated and treated samples was determined by X-ray
diffraction (XRD) using a Bruker D8 Advance diffractometer (Bruker AXS, Berlin, Germany) with
standard monochromatic Cu–Kα radiation at 40 kV and 30 mA with a Ni filter and Linxeye 1D detector.
Routine scanning was performed with a 0.015◦ 2θ step size, and at 0.1 s per step from 3◦ to 70◦ .
Rietveld refinement was also realized to determine the quantitative composition of the untreated
bricks. In this case, scanning was performed with a 0.010◦ 2θ step size at 0.5 s per step in the range of
3◦ –120◦ and adding zincite (15 wt %) as an internal standard. Rietveld refinement for the present phase
quantification was done with Bruker’s commercial Topas v5 software (Bruker AXS, Berlin, Germany).
In addition, previous carbonation analysis was carried out by X-ray diffraction using a Bruker D8
Advance powder diffractometer (Bruker AXS, Berlin, Germany) equipped with an Anton Paar XRK
900 reactor chamber (Anton Parr GmbH, Graz, Austria) and high-sensitivity detector Bruker Vantec 1
(Bruker AXS, Berlin, Germany). This chamber was designed for X-ray diffraction experiments of up to
900 ◦ C and 10 bar for solid state–gas reactions. Standard monochromatic Cu–Kα radiation operating at
40 kV and 40 mA was employed. Scanning was performed with a 0.022◦ 2θ step size at 0.2 s per step
from 3◦ to 70◦ . Samples were in a CO2 -rich environment for 24 h in this reactor chamber.
Macro- and micro-observations were obtained by stereomicroscope using a Greenough
Leica S8 APO (Leica Microsystems GmbH, Wetzlar, Germany) equipped with a DC300 camera
(Leica Microsystems GmbH, Wetzlar, Germany) and by scanning electronic microscopy (SEM) using a
JEOL 6460 LV microscope (JEOL Ltd., Akishima, Japan) equipped with energy-dispersive spectrometers
(Oxford Instruments INCA, Oxford, UK).
The carbonate content of the carbonated samples was determined by two analytical methods:
differential thermal and thermogravimetric analysis (DTA-TG) and an elemental analyzer. DTA-TG
were performed on a TG Netzsch STA 409 PC. Samples (around 150 mg) were heated in an aluminum
oxide crucible under a nitrogen atmosphere at 10 ◦ C min−1 from room temperature to 1200 ◦ C.
Weight loss was measured by thermogravimetric analysis in the temperature range of 450–900 ◦ C
relative to the total carbonated decomposition. Elemental carbon content was measured using an
elemental analyzer, Leco Truspec CHNS Micro (St. Joseph, MI, USA), which calculated the carbonated
ratio by assuming that the whole carbon content was calcite.
Soluble Si, Ca, and Mg ions of the original and treated samples were also determined. Analysis was
performed with simultaneous inductively coupled plasma optical emission spectrometry (IPC-OES)
analysis using a Horiba Jobin Yvon ULTIMA 2 model instrument (Horiba Scientific, Palaiseau, France).
The samples were prepared by mixing the solid-powder samples with water and stirring for 24 h,
isolating the liquid phase by centrifugation, and filtering using a Nylon 0.22 µm syringe filter
(MilliporeSigma, Burlington, MA, USA).
Specific surface area (BET) and microporosity were measured with a Micromeritics Gemini
2360 instrument (Micromeritics Instrument Corp, Norcross, GA, USA) using the absorption of N2 at
liquid nitrogen temperature. Before measuring, all samples were degassed using a Flow Prep 060
Micromeritics degasser (Micromeritics Instrument Corp, Norcross, GA, USA) with dry nitrogen gas at
80 ◦ C for 12 h. Nanoporosity was measured with an ASAP 2420 instrument (Micromeritics Instrument
Corp, Norcross, GA, USA) using CO2 absorption at room temperature. Samples were degassed at
150 ◦ C for 1.5 h and finally outgassed to 10−3 Torr. Macro- and mesoporosity were studied using
mercury porosimeter Quantachrome Instruments Pore Master 60-GT (Quantachrome Instruments,
Boynton Beach, FL, USA).
Minerals 2019, 9, 612 4 of 14

3. Results and Discussion

3.1. Proposed Methodological Sequence to Evaluate Effectiveness of Construction and Demolition Waste for CCS
A first outcome measure to take into consideration to provide the optimal construction residue is
to have an important content of calcium oxide and/or magnesium as reactive compounds, since these
elements are necessary for precipitation in the form of the carbonate, although the carbonation of other
alkali metals is possible, as shown in the following equations [25]:

MO + CO2 → MCO3 . (1)

This reaction is usually exothermic in nature, as per Lackner et al. (1995) [26]:

CaO + CO2 → CaCO3 + 179 kJ/mole (2)

MgO + CO2 → MgCO3 + 118 kJ/mole. (3)

Consequently, sample characterization requires a chemical analysis that is commonly used (WDXRF).
It has been widely described in the literature that the main possible carbonation minerals are
oxides, silicates, and anhydrite; therefore, it was necessary to determine the minerals present in
the sample in order to evaluate the candidate. Natural wollastonite is a widely studied mineral in
several works, for example, Huijgen and coworkers [27], as a candidate for mineral carbonation.
Different studies used other types of sample, such as industrial waste, that were mainly composed of
wollastonite [28–30].
The proposed methodology includes a sample preselection using the characterization of chemical
and mineralogical conditions, and in situ carbonation. The second part of the methodology is
carbonation tests in the selected samples. The flow diagram of the proposed methodology is shown in
Figure 1.

Figure 1. Flow diagram of proposed methodology.

For that reason, mineral characterization by powder X-ray diffraction is commonly used. In the
same way, we propose an in situ carbonation test and process the evaluation by using a diffractometer
equipped with a reaction chamber in a CO2 -rich environment. We mixed 2 g of the powdered sample
and a couple of pipetted drops, and took a scan every six hours over the course of 24 h. These tests
Minerals 2019, 9, 612 5 of 14

made it possible to obtain the first analysis of carbonation evolution, selecting a feasible candidate for
the carbonation tests.
Once the carbonation tests were completed, the treated samples had to be analyzed by XRD,
showing the presence of carbonate phases as carbonation by-products. Rietveld refinement of both
the treated and the original sample allowed the quantification of the present phases, determining the
percentage of new stable carbonate phases and the destruction (or partial destruction) of phases that
provided the necessary calcium and/or magnesium.
Likewise, the presence of new precipitated minerals could be observed by SEM. Combined with
microchemical composition from energy-dispersive spectrometers (EDS), it confirmed the possible
by-product phase. If the precipitate was large enough, then the optical microscopy supply provided
information on how the precipitate grew on the surface of the sample and texture, as well as
estimating sizes.
As an alternative to the quantification processes, and especially when these new phases did
not represent a sufficient percentage for correct quantification, to quantify the amount of CO2
capture, two techniques were employed: elemental carbon measure and weight loss by differential
thermal and thermogravimetric analysis (DTA-TG) in the temperature range of carbonate mineral
decomposition. In the first case, CO2 capture can be calculated by the direct conversion of carbon to
dioxide (CO2 (wt %) = 3.6641 × C (wt %)) for the difference of the carbon content in the original and the
treated sample. In the second one, weight-loss measuring in the right temperature range corresponding
to the thermal decomposition of carbonates for the difference between treated and the original sample
corresponded to the percentage by weight of captured CO2 . Since it was different, if any of the minerals
constituting the samples had total or partial decomposition in the same temperature range, they would
not be quantified, since they would be present in both samples pre- and post-treatment, for example,
carbonates that already originally existed. The evolution of soluble ion content helps what follows
mineral destruction.
Another aspect to take into consideration was the evaluation of the specific surface area by
Brunauer–Emmett–Teller (BET) analysis. This technique can help us understand how the specific
surface of a sample evolves at the microscopic level (the texture). Complete porosity analysis from
the nano- to the macroscale also allows a study of the porous system after carbonation by isotherm
adsorption of CO2 and N2 . In this sense, it was possible to analyze how the precipitate completed the
sample-surface pores, reducing their size to a smaller scale from the macro- or mesoscale to the micro-
or nanoscale.
Finally, it is common in the literature to use the Steinour formula [21,24,31–33] to determine the
efficiency of the reaction from the theoretical maximal CO2 sequestration value obtained from the
following stoichiometric formula (Equation (4)):

CO2 (wt %) = 0.785(%CaO − 0.7%SO3 )+1.09%MgO + 0.71%Na2 O + 0.468%K2 O. (4)

3.2. Validation of Proposed Methodology Using Ca-Silicate-Rich Brick


For each of the techniques described in the previous section, an example of the obtained results for
the selected sample for this study (MPC2) and the correlations between them is presented. However,
this example sample was already studied in more detail in a previous work of some of the authors in
mineral carbonation processes [5].
In the MPC2 sample, CaO was around 15 wt % and 2 wt % for MgO (Table 1 and Tables S1–S3
in Supplementary Materials), which were appropriate values to be considered candidates for the
mineral-carbonation process. This sample was composed of quartz, k- and alkali feldspar, plagioclases
(orthoclase, anorthite, and albite), and calcium-rich silicate as wollastonite (Figure 2 and Table 2).
Minerals 2019, 9, 612 6 of 14

Table 1. Chemical composition of original brick MPC2 (wt %) by X-ray fluorescence (XRF) (modified
from [5]).

Sample SiO2 Al2 O3 Fe2 O3 MnO MgO CaO Na2 O K2 O TiO2 P2 O5 SO3 LOI TOTAL
MPC2 56.8 16.6 5.4 0.1 1.9 14.8 0.6 2.6 0.9 0.1 0.3 0.8 100.8
Detection Limit (DL) 0.01 0.01 0.01 0.02 0.01 0.04 0.01 0.02 0.03 0.01 0.22
Quantification Limit (QL) 0.02 0.02 0.02 0.03 0.02 0.05 0.03 0.03 0.10 0.02 0.23
Relative Error 0.012 0.020 0.058 0.184 0.007 0.047 0.038 0.028 0.061 0.025 0.063
LOI: loss on ignition at 1025 ◦ C.

Figure 2. X-ray diffraction (XRD) pattern of MPC2 (and Rietveld refinement). Abbrv.: Qtz—quartz;
Ab—Albite; Wo—wollastonite; Or—orthoclase; Di—diopside; Zi—zincite (Internal Standard).

Table 2. Mineralogical composition (wt %) of selected brick MPC2 (Qtz: quartz; Wo: wollastonite; Or:
orthoclase; Ab: albite; Di: diopside; An: anorthite; Amor: amorphous phase). Rexp, Rwp, and GOF
are numerical indicators of how well the Rietveld model was refined. Rwp, residual of least-squares
refinement (weighted), which must be improved in refinement (with common sense); Rexp evaluates
data quality; and GOF, goodness of fit parameter.

Sample Qtz Wo Or Ab Di An Amor


MPC2 20.5 6.2 3.7 11.8 4.7 29.5 23.6
Rietveld Refinement Rexp: 1.86 GOF: 1.72 Rwp: 3.21

An in situ carbonation test (Figure 3) showed the evolution of newly grown calcite over time
in a CO2 -rich environment. Intensity for the main calcite’s peak (at d = 3.04 Å) increased directly
proportional to time.
The next step was to perform carbonation tests on a laboratory scale with the selected bricks according
to previous analysis, the high content of CaO, mineralogical composition rich in Ca silicate, and the
presence of calcite in the in situ carbonation test. Tests were performed at room temperature and 10 bar
pressure for different reaction times, and three fractions of particle sizes were representative of the total.
Additionally, a test was carried out for the original sample at low pressure (1 bar), room temperature,
and a 4:1 solid–water ratio in a 5 L volume hermetic reactor for 720 h of reaction time.
In the X-ray patterns of treated samples, compared with the nontreated, the most obvious
differences were the presence of calcite, and the partial destruction of wollastonite and some orthoclase.
The attack on wollastonite with carbonic acid allowed for the release of calcium ions and calcite
precipitation [4–6,27,28,34–36].
Minerals 2019, 9, 612 7 of 14
Minerals 2019, 9, x FOR PEER REVIEW 7 of 15

Figure
Figure In situ
3. In3.situ XRD XRD MPC2
MPC2 patternwhere
pattern where d
d== 3.04
3.04 Å,
Å , reflection
reflectioncorresponding
correspondingto precipitated newly
to precipitated newly
formed calcite.
formed calcite.
Therefore, the carbonation process occurred in the following two steps: a) silicate mineral
The next step was to perform carbonation tests on a laboratory scale with the selected bricks
dissolution and b) carbonate precipitation. Several studies based on the mineral carbonation of
according to previous
wollastonite [27,28] analysis,
described the
the high content
process of CaO,carbonation
in an aqueous mineralogical composition
route as: rich in Ca silicate,
and the presence of calcite in the in situ carbonation test. Tests were performed at room temperature
and 1.10 Dissolution
bar pressure of CO 2 indifferent
for water for the production
reaction of aand
times, (bi-)carbonate:
three fractions of particle sizes were
representative of the total. Additionally, a test was carried out for− the original sample at low pressure
CO2 (g) + H2 O (l) → H2 CO3 (aq) → HCO3 (aq) + H+ (aq). (5)
(1 bar), room temperature, and a 4:1 solid–water ratio in a 5 L volume hermetic reactor for 720 h of
reaction
2. time.
Leaching Ca from wollastonite by acidic attack:
In the X-ray patterns of treated samples, compared with the nontreated, the most obvious
differences were the presence CaSiOof + 2H+ (aq)
3 (s)calcite, and→theCa2+ (aq) + H
partial 2 O (l) + SiO2of
destruction (s).wollastonite and(6)some
orthoclase. The attack on wollastonite with carbonic acid allowed for the release of calcium ions and
3. Nucleation and growth of calcium carbonate:
calcite precipitation [4–6,28,29,35–37].
Therefore, the carbonation process occurred in the following two steps: a) silicate mineral
Ca2+ (aq) + HCO3 − (aq) → CaCO3 (s) + H+ (aq). (7)
dissolution and b) carbonate precipitation. Several studies based on the mineral carbonation of
wollastonite [28,29] described the process in an aqueous carbonation route as:
These new carbonate crystals were observed by stereomicroscope (optical microscopy) and
1. Dissolution
scanning of CO2 in(Figures
electron microscopy water for the5),production
4 and of a (bi-)carbonate:
with a composition close to theoretical calcite, as shown
by the results of elemental analysis by EDS.
CO2 (g) + H2O (l)  H2CO3 (aq)  HCO3− (aq) + H+ (aq). (5)

2. Leaching Ca from wollastonite by acidic attack:

CaSiO3 (s) + 2H+ (aq)  Ca2+ (aq) + H2O (l) + SiO2 (s). (6)

3. Nucleation and growth of calcium carbonate:

Ca2+ (aq) + HCO3− (aq)  CaCO3 (s) + H+ (aq). (7)

These new carbonate crystals were observed by stereomicroscope (optical microscopy) and
scanning electron microscopy (Figures 4 and 5), with a composition close to theoretical calcite, as
shown by the results of elemental analysis by EDS.
Minerals 2019,
Minerals 9, x612
2019, 9, FOR PEER REVIEW 88of
of 15
14

4. Layers
Figure 4.
Figure Layersofofcalcite onon
calcite thethe
MPC2 surface,
MPC2 observed
surface, by stereomicroscope
observed after 720after
by stereomicroscope h of CO
7202 treatment.
h of CO2
treatment.

Figure 5. Scanning-electron-microscopy (SEM) image, and the energy-dispersive-spectrometer (EDS)


spectrum and elemental quantification of MPC2.
Figure 5. Scanning-electron-microscopy (SEM) image, and the energy-dispersive-spectrometer (EDS)
spectrum and elemental
The measurement quantification
of elemental of MPC2.
carbon and weight loss by DTA-TG in the temperature range of
the calcite decomposition (Table 3 or Figure 6a) was used to quantify the amount of captured CO2 .
The measurement of elemental carbon and weight loss by DTA-TG in the temperature range of
Both results could be expressed in terms of calcite as carbonate ore. Obviously, it was an indirect
the calcite decomposition (Table 3 or Figure 6a) was used to quantify the amount of captured CO2.
method due to it being assumed, on the one hand, that the only phase that decomposes in that
Both results could be expressed in terms of calcite as carbonate ore. Obviously, it was an indirect
temperature range was calcite and, on the other hand, that it only precipitates calcite as a carbonated
method due to it being assumed, on the one hand, that the only phase that decomposes in that
material. This weight loss also corresponded to the expulsion of CO2 and carbon content that was
temperature range was calcite and, on the other hand, that it only precipitates calcite as a carbonated
related to the theoretical content of CO2 in the chemical composition of calcite (Equations (8) and
material. This weight loss also corresponded to the expulsion of CO2 and carbon content that was
(9)) [37,38].
related to the theoretical content of CO2 in the chemical composition of calcite (Equations (8) and (9))
[37,38]. ∆m ∆m450−900°C
% CalciteDTA−TG = CO 450−900°C
Theorical × 100 = 43.97 × 100 = 2.274 ×
2∆ ℃ ∆ ℃ (8)
% Calcite = ×
∆m 100 =
∆m450−900°C % CalciteDTA−TG = CO Theorical × 100
450−900°C × 450−900°C
∆m 100 = 2.274 × 2.274 × ∆m
. = 43.97 × 100 = 450−900°C
2

∆ ∆
∆m % Calcite = Ccontent ℃ × 100 =Ccontent ℃ × 100 25
= 2.274 × ∆m
%℃CalciteC−Elmental = × 100 = . × 100 = Ccontent ℃ (9)
CTheorical 12 3
(8)

% 𝐶𝑎𝑙𝑐𝑖𝑡𝑒 = × 100 = × 100 = 𝐶 . (9)


Minerals 2019, 9, 612 9 of 14
Minerals 2019, 9, x FOR PEER REVIEW 10 of 15

Figure 6. (a) MPC2 thermogravimetry (solid line) and differential thermal analyses (dashed line);
>4 mm6.and
Figure (a) 720
MPC2h reaction time. (b) Calculated
thermogravimetry calcite
(solid line) content by DTA-TG
and differential (circle) and
thermal analyses by C line);
(dashed element>4
(square) as function of reaction time and particle size fraction (modified from [5]).
mm and 720 h reaction time. (b) Calculated calcite content by DTA-TG (circle) and by C element
(square) as function of reaction time and particle size fraction (modified from [5]).
Minerals 2019, 9, 612 10 of 14

Table 3. Carbon elemental content, mass loss by differential thermal and thermogravimetric analysis
(DTA-TG), and calcite content calculated by both techniques.

Reaction
Particle Size C-Elemental Mass Loss Calcite Calcite
Time
(mm) (hours) (wt %) (wt %) (wt %) (wt %)
MPC2 by C-Elemental by DTA-TG
Original 0 0.08 0.15 0.63 0.34
>4 mm 24 0.10 0.20 0.87 0.45
>4 mm 120 0.21 0.61 1.74 1.39
>4 mm 240 0.36 1.49 3.02 3.39
>4 mm 720 0.73 2.80 6.10 6.37
2–4 mm 24 0.30 0.61 2.51 1.39
2–4 mm 120 0.39 1.10 3.26 2.50
2–4 mm 240 0.53 1.99 4.38 4.53
2–4 mm 720 0.80 3.30 6.67 7.51
1–2 mm 24 0.21 0.67 1.79 1.52
1–2 mm 120 0.38 1.26 3.19 2.87
1–2 mm 240 0.63 1.75 5.29 3.98
1–2 mm 720 0.77 3.33 6.38 7.57

The difference compared to carbon analysis using the elemental analyzer could be attributed to
the sum of experiment errors and analysis sensitivity. Yet, it was also necessary to take account of the
adsorbed CO2 , which could be measured in CHNS instead of DTA. However, both followed the same
tendency with respect to reaction times (Figure 6b). In both instances, calcite content was higher than
the original and directly proportional to the reaction time in the studied size fractions.
The wollastonite attack with carbonic acid allowed for the release of calcium ions and calcite
precipitation. Concerning the presence of soluble ions (Figure 7 and Table S4 in Supplementary
Materials), the amount of Ca ions decreased over reaction time because of calcite precipitation, while Si
increased as a consequence of partial silicate destruction. This partial destruction of silicates resulted
in a new specific surface on the bulk (Figure 8). The increase of the specific surface had a direct
relationship reaction time. The newly precipitated calcite on the sample surface also increased BET.
However, both possibilities must have had a greater effect than physical CO2 absorption that would
result in a reduction of the specific surface area.

Figure 7. Soluble Ca, Mg, and Si ions measured untreated and treated after 240 and 720 h of reaction
for MPC2 (modified from [5]).

After the carbonation tests, the brick samples revealed macro- and mesoporosity as determined
by Hg porosity, which affects the proportion and size of the pores (Figure 9a,b), with a decrease in
microporosity and increase in nanoporosity by N2 and CO2 absorption (Figure 9c,d). All of them were
a result of two differentiated processes: (a) the action of carbonic acid destroying calcium silicates that
produced an irregular surface and an increment of macro- and mesoporosity, and (b) the precipitation
of carbonates that filled the micropores and probably reduced them to nanopores.
Minerals 2019, 9, 612 11 of 14

Figure 8. Specific surface evolution by specific surface area (BET) as a function of reaction time for
MPC2 (modified from [5]).

Figure 9. (a,b) Histograms of Hg porosity, (c) N2 adsorption, and (d) CO2 adsorption isotherms for
original brick and treated brick layer.
Minerals 2019, 9, 612 12 of 14

Calculating efficiency was as the quotient between the percentage of captured CO2 (experimental)
and theoretical CO2 by Steinour’s formulae (Equation (4); Table 4). The calculated capture efficiency
was proportional to the reaction time (the longer the time was, the higher the carbonation amount),
whereas carbonation did not seem to significantly depend on particle size in the studied conditions.
Results obtained at low pressure (1 bar) after 720 h of reactions for the total sample were very similar
to those obtained for the finer fractions (2–4, 1–2 mm) at 10 bars.

Table 4. Carbonation efficiency (CE) according to Steinour’s equation.

Particle Size Pressure Reaction Time CO2 Exp Efficiency


(mm) (bar) (hours) (wt %) (%)
MPC2
Original 10 0 0.15 1.00
>4 mm 10 24 0.20 1.32
>4 mm 10 120 0.61 4.08
>4 mm 10 240 1.49 9.94
>4 mm 10 720 2.80 18.67
2–4 mm 10 24 0.61 4.08
2–4 mm 10 120 1.10 7.33
2–4 mm 10 240 1.99 13.28
2–4 mm 10 720 3.30 22.02
1–2 mm 10 24 0.67 4.46
1–2 mm 10 120 1.26 8.41
1–2 mm 10 240 1.75 11.67
1–2 mm 10 720 3.33 22.19
Total Fraction 1 720 3.20 21.33

4. Conclusions
The present research showed a suitable methodology to evaluate the possibility of using ceramic
construction waste (or other types of construction and demolition waste) for mineral carbonation
under surface conditions in short periods of time. Even though the experiment was conducted in
comparatively large scale, each test method yielded only a very small fraction. Therefore, to increase
reliability each test could be conducted several times.
Specifically, Ca-rich bricks were successfully used as raw material for direct mineral carbonation
by the destruction of Ca silicates. The amount of carbonation was proportional to the reaction time,
whereas it did not seem to significantly depend on particle size or pressure in the studied conditions.
Acceptable carbonation efficiency was achieved under the favorable conditions of low pressure
and temperature.
These results open the possibility for future studies using the proposed methodology for other
types of construction and demolition waste rich in calcium silicates or other calcium compounds,
which could be directly carbonated by in situ injections of CO2 .

Supplementary Materials: The following are available online at http://www.mdpi.com/2075-163X/9/10/612/s1, Table


S1: List of certificated standars used for XRF calibration.; Table S2: Detection limit (L.D.), quantification limit
(L.C.) and relative error for the measuremento of the mayor elements by XRF in Panalytical Axios spectrometer of
the SGI Laboratorio de Rayos X (University of Seville) [September 2014]; Table S3: Mayor elements comparative
between monitor standards (calibration validation) in white box and the certificate value in grey box [September
2014]; Table S4: Soluble Ca, Mg, and Si ions measured untreated and treated after 240 and 720 h of reaction for
MPC2 and their standard deviations.
Author Contributions: D.M. collected and analyzed the data and wrote the paper; V.F.-A. conceived and designed
the ideas; and revised the paper; P.A. made the formal Investigation, discuss the data and revised/edited
the manuscript.
Minerals 2019, 9, 612 13 of 14

Funding: This research was funded by the Junta de Andalucía (Consejería de Economía y Conocimiento)
(P12-RNM-568 MO project) and the contract of Domingo Martín granted by the V Plan Propio de Investigación de
la Universidad de Sevilla.
Acknowledgments: Authors are grateful to editor and reviewers for their comments which improved the
manuscript. XRD, XRF, ICP-OES, C-elemental and SEM analysis were performed using the facilities of the General
Research Center at the University of Seville (CITIUS).
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Alberola, E.; Chevallier, J.; Chèze, B. The EU emissions trading scheme: The effects of industrial production
and CO2 emissions on carbon prices. Econ. Int. 2008, 4, 93–125. [CrossRef]
2. Caldeira, K.; Jain, A.K.; Hoffert, M.I. Climate sensitivity uncertainty and the need for energy without CO2
emission. Science 2003, 299, 2052–2054. [CrossRef] [PubMed]
3. Pan, S.-Y.Y.; Chang, E.E.; Chiang, P.C. CO2 capture by accelerated carbonation of alkaline wastes: A review
on its principles and applications. Aerosol Air Qual. Res. 2012, 12, 770–791. [CrossRef]
4. Martín, D.; Aparicio, P.; Galán, E. Mineral carbonation of ceramic brick at low pressure and room temperature.
A simulation study for a superficial CO2 store using a common clay as sealing material. Appl. Clay Sci.
2018, 161, 119–126. [CrossRef]
5. Martín, D.; Aparicio, P.; Galán, E. Accelerated Carbonation of Ceramic Materials. Application to Bricks from
Andalusian Factories (Spain). Constr. Build. Mater. 2018, 181, 598–608. [CrossRef]
6. Martín, D.; Aparicio, P.; Galán, E. Time evolution of the mineral carbonation of ceramic bricks in a simulated
pilot plant using a common clay as sealing material at superficial conditions. Appl. Clay Sci. 2019, 180, 105191.
[CrossRef]
7. Eighmy, T.; Gardner, K.; Seager, T. Method for Sequestering Carbon Dioxide. U.S. Patent US20050238563A1,
27 October 2005. Available online: https://patentimages.storage.googleapis.com/bb/49/8c/56d81ea0f8406a/
US20050238563A1.pdf (accessed on 3 October 2019).
8. Montes-Hernandez, G.; Pérez-López, R.; Renard, F.; Charlet, L.; Nieto, J.-M. Process for Sequestration of CO2
by Reaction with Alkaline Solid Waste. European Patent EP 07 1233005.5, 14 December 2007. Available online:
https://patentimages.storage.googleapis.com/47/a0/49/787d3e0e245078/WO2007071633A1.pdf (accessed on
3 October 2019).
9. Mayoral, M.C.; Andrés, J.M.; Gimeno, M.P. Optimization of mineral carbonation process for CO2 sequestration
by lime-rich coal ashes. Fuel 2013, 106, 448–454. [CrossRef]
10. Santos, R.M.; Van Bouwel, J.; Vandevelde, E.; Mertens, G.; Elsen, J.; Van Gerven, T. Accelerated mineral
carbonation of stainless steel slags for CO2 storage and waste valorization: Effect of process parameters on
geochemical properties. Int. J. Greenh. Gas Control 2013, 17, 32–45. [CrossRef]
11. Sipilä, J.; Teir, S.; Zevenhoven, R. Carbon Dioxide Sequestration by Mineral Carbonation Literature Review
Update 2005–2007. Report VT 2008-1. p. 52. Available online: http://innovationconcepts.nl/res/literatuurGPV/
mineralcarbonationliteraturereview0507.pdf (accessed on 3 October 2019).
12. Soong, Y.; Fauth, D.L.; Howard, B.H.; Jones, J.R.; Harrison, D.K.; Goodman, A.L.; Gray, M.L.; Frommell, E.A.
CO2 sequestration with brine solution and fly ashes. Energy Convers. Manag. 2006, 47, 1676–1685. [CrossRef]
13. Uibua, M.; Kuusik, R.; Andreas, L.; Kirsimäe, K. The CO2 -binding by Ca-Mg-silicates in direct aqueous
carbonation of oil shale ash and steel slag. Energy Procedia 2011, 4, 925–932. [CrossRef]
14. Gunning, P.J.; Hills, C.D.; Carey, P.J. Accelerated carbonation treatment of industrial wastes. Waste Manag.
2010, 30, 1081–1090. [CrossRef] [PubMed]
15. Geerlings, J.J.C.; Van Mossel, G.A.F.; Veen, B.C.M.I. Process for Sequestration of Carbon Dioxide by Mineral
Carbonation. U.S. Patent US20100196235A1, 25 May 2010. Available online: https://patentimages.storage.
googleapis.com/47/b3/78/16189676247184/US20100196235A1.pdf (accessed on 3 October 2019).
16. Kawatra, S.K.; Eisele, T.C.; Simmons, J.J. Capture and Sequestration of Carbon Dioxide in Flue Gases. U.S.
Patent 7.919.064B2, 5 April 2011. Available online: https://patentimages.storage.googleapis.com/3d/6c/86/
beea6a2fc7f1ba/US7919064.pdf (accessed on 3 October 2019).
Minerals 2019, 9, 612 14 of 14

17. Riman, R.E.; Atakan, V. Systems and Methods for Carbon Capture and Sequestration and Compositions
Derived Therefrom. U.S. Patent US8114367B2, 14 February 2012. Available online: https://patentimages.
storage.googleapis.com/72/67/38/c293969a42ecee/US8114367.pdf (accessed on 3 October 2019).
18. Galán, E.; Aparicio, P. Captura y Secuestro de CO2 Mediante la Carbonatición de Residuos Cerámicos.
WITO Patent WO2011089292A1, 28 July 2011.
19. Costa, G.; Baciocchi, R.; Polettini, A.; Pomi, R.; Hills, C.D.; Carey, P.J. Current status and perspectives
of accelerated carbonation processes on municipal waste combustion residues. Environ. Monit. Assess.
2007, 135, 55–75. [CrossRef] [PubMed]
20. Dri, M.; Sanna, A.; Maroto-Valer, M.M. Mineral carbonation from metal wastes: Effect of solid to liquid ratio
on the efficiency and characterization of carbonated products. Appl. Energy 2014. [CrossRef]
21. Huntzinger, D.N.; Gierke, J.S.; Sutter, L.L.; Kawatra, S.K.; Eisele, T.C. Mineral carbonation for carbon
sequestration in cement kiln dust from waste piles. J. Hazard. Mater. 2009, 168, 31–37. [CrossRef]
22. Li, X.; Bertos, M.F.; Hills, C.D.; Carey, P.J.; Simon, S. Accelerated carbonation of municipal solid waste
incineration fly ashes. Waste Manag. 2007, 27, 1200–1206. [CrossRef]
23. Haselbach, L. Potential for carbon dioxide absorption in concrete. J. Environ. Eng. 2009, 135, 465–472.
[CrossRef]
24. Fernández Bertos, M.; Simons, S.J.R.; Hills, C.D.; Carey, P.J. A review of accelerated carbonation technology in the
treatment of cement-based materials and sequestration of CO2 . J. Hazard. Mater. 2004, 112, 193–205. [CrossRef]
25. Seifritz, W. CO2 disposal by means of silicates. Nature 1990, 345, 486. [CrossRef]
26. Lackner, K.S.; Wendt, C.H.; Butt, D.P.; Joyce, E.L.; Sharp, D.H. Carbon dioxide disposal in carbonate minerals.
Energy 1995, 20, 1153–1170. [CrossRef]
27. Huijgen, W.J.J.; Witkamp, G.J.; Comans, R.N.J. Mechanisms of aqueous wollastonite carbonation as a possible
CO2 sequestration process. Chem. Eng. Sci. 2006, 61, 4242–4251. [CrossRef]
28. Daval, D.; Martinez, I.; Corvisier, J.; Findling, N.; Goffé, B.; Guyot, F. Carbonation of Ca-bearing silicates,
the case of wollastonite: Experimental investigations and kinetic modeling. Chem. Geol. 2009, 262, 262–277.
[CrossRef]
29. Huijgen, W.J.J.; Ruijg, G.J.; Comans, R.N.J.; Witkamp, G.J. Energy consumption and net CO2 sequestration of
aqueous mineral carbonation. Ind. Eng. Chem. Res. 2006, 45, 9184–9194. [CrossRef]
30. Tai, C.Y.; Chen, W.; Shih, S. Factors affecting wollastonite carbonation under CO2 supercritical conditions.
AIChE J. 2006, 52, 292–299. [CrossRef]
31. Huntzinger, D.N.; Gierke, J.S.; Kawatra, S.K.; Eisele, T.C.; Sutter, L.L. Carbon Dioxide Sequestration in Cement
Kiln Dust through Mineral Carbonation. Environ. Sci. Technol. 2009, 43, 1986–1992. [CrossRef] [PubMed]
32. Steinour, H.H. Some effects of carbon dioxide on mortars and concrete-discussion. J. Am. Concr. Inst
1959, 30, 905–907.
33. Yixin, S.; Xudong, Z.; Monkman, S. A new CO2 sequestration process via concrete products production.
In Proceedings of the 2006 IEEE EIC Climate Change Conference, Ottawa, ON, Canada, 10–12 May 2006.
34. Wang, W.; Xiao, J.; Wei, X.; Ding, J.; Wang, X.; Song, C. Development of a new clay supported polyethylenimine
composite for CO2 capture. Appl. Energy 2014, 113, 334–341. [CrossRef]
35. Mun, M.; Cho, H. Mineral carbonation for carbon sequestration with industrial waste. Energy Procedia
2013, 37, 6999–7005. [CrossRef]
36. Prigiobbe, V.; Hänchen, M.; Werner, M.; Baciocchi, R.; Mazzotti, M. Mineral carbonation process for CO2
sequestration. Energy Procedia 2009, 1, 4885–4890. [CrossRef]
37. Lim, M.; Han, G.C.; Ahn, J.W.; You, K.S. Environmental remediation and conversion of carbon dioxide
(CO2 ) into useful green products by accelerated carbonation technology. Int. J. Environ. Res. Public Health
2010, 7, 203–228. [CrossRef]
38. Villagrán-Zaccardi, Y.A.; Egüez-Alava, H.; De Buysser, K.; Gruyaert, E.; De Belie, N. Calibrated quantitative
thermogravimetric analysis for the determination of portlandite and calcite content in hydrated cementitious
systems. Mater. Struct. 2017, 50, 179. [CrossRef]

© 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like